Abstract
In the past several years, the relationship between chromatin structure and mRNA processing has been the source of significant investigation across diverse disciplines. Central to these efforts was an unanticipated nonrandom distribution of chromatin marks across transcribed regions of protein-coding genes. In addition to the presence of specific histone modifications at the 5′ and 3′ ends of genes, exonic DNA was demonstrated to present a distinct chromatin landscape relative to intronic DNA. As splicing in higher eukaryotes predominantly occurs co-transcriptionally, these studies raised the possibility that chromatin modifications may aid the spliceosome in the detection of exons amidst vast stretches of noncoding intronic sequences. Recent investigations have supported a direct role for chromatin in splicing regulation and have suggested an intriguing role for splicing in the establishment of chromatin modifications. Here we will summarize an accumulating body of data that begins to reveal extensive coupling between chromatin structure and pre-mRNA splicing.
Keywords: Alternative splicing, Chromatin, RNA polymerase II, Transcription, Epigenetics
1. Introduction
Unlike the genes of lower eukaryotes, in which protein-coding sequences are typically uninterrupted, genes of higher metazoans are characterized by a large number of coding exons separated by long stretches of noncoding introns. As genes are transcribed into mRNA, introns are excised by the megadalton spliceosome complex, which recognizes short consensus sequences at intron–exon boundaries [1]. While introns were initially dubbed as “junk DNA,” it is increasingly evident that exon–intron architecture serves as a critical platform for transcriptome diversification via alternative pre-mRNA splicing. Cassette exons thus represent an important aspect of proteome complexity in higher organisms, and current estimates indicate that greater than 90 % of human genes engage in alternative splicing [2, 3]. However, the evolutionary drive for transcriptome expansion has posed the spliceosome with an increasingly difficult task as intron lengths have increased and splice site strengths have weakened [4]. Alternative pre-mRNA splicing adds an additional layer of complexity in that splice site recognition must be diversified in a context-dependent manner. To accomplish regulated transcript production within a multivariable framework, pre-mRNA splicing is coordinated at multiple levels. In addition to regulation via RNA-binding protein recognition of cis-elements encoded within pre-mRNA [5], the rate of RNA polymerase II (pol II) transcription elongation and chromatin structure contribute to splice site recognition [6]. Rather than operating independently, these processes are highly integrated as a result of co-transcriptional pre-mRNA splicing [7, 8]. The basic mechanisms of alternative splicing regulation via RNA-binding proteins and evidence for co-transcriptional splicing are discussed elsewhere in this volume. Here, we will focus directly on the accumulating evidence in support of a role for chromatin structure in splicing regulation.
2. Chromatin and Co-transcriptional Splicing
A central tenet to the relationship between chromatin structure and alternative splicing is that the majority of splicing in higher eukaryotes occurs co-transcriptionally, while the nascent message is still tethered to the template DNA. This allows for several layers of coupling between the transcription and splicing machineries. Initial efforts to address coupling focused on the potential association between RNA-binding proteins and the pol II carboxy-terminal domain (CTD), such that the factors could be efficiently transferred to the nascent transcript during the process of transcription [9]. Combinatorial association of these factors could in principle influence splicing decisions [5]. Co-transcriptionality further allows for kinetic regulation of splicing decisions. In work pioneered in the Kornblihtt group, it was shown that the rate of transcription elongation impacts splicing decisions such that weak exons are more likely to be excluded from spliced mRNA in response to a rapid elongation rate [10]. While the kinetic model has now been validated in a variety of systems, the physiological barriers to pol II elongation remained comparatively elusive until recently. Genome-wide profiling of chromatin modifications revealed the transcribed DNA template itself as a potential modulator of elongation rate or other aspects of splicing regulation. Intragenic DNA presents a distinct chromatin landscape relative to intergenic DNA, and more importantly to this discussion, exonic DNA presents unique features relative to intronic DNA [11]. These observations raised the possibility that the chromatin structure of transcribed genes may aid the spliceosome in the process of exon definition. In this section, we expand on these themes and examine the various evidences and mechanisms for chromatin-directed splicing.
2.1. Evidence for Co-transcriptional Splicing
While pre-mRNA splicing was initially envisioned as a distinct cellular process that occurred subsequent to the completion of transcription, evidence for co-transcriptional splicing quickly mounted. In a widely cited landmark study, electron micrographs of chromosomal spreads from Drosophila embryos provided a visual demonstration of spliceosome assembly on nascent mRNA, while the RNA was still tethered to the template DNA [12, 13]. Further, indirect evidence of functional cross talk between transcription and pre-mRNA splicing came from studies wherein protein-coding genes were placed downstream of RNA polymerase I or RNA polymerase III promoters [14–17]. These studies indicated that pol I and pol III promoters led to efficient mRNA transcription, but the resulting RNAs were poorly spliced, highlighting an obligatory functional connection between pol II and the splicing machinery [14–17]. Similar results were obtained following deletion of the pol II CTD [18]. The CTD consists of multiple repeats (52 in human) of evolutionary conserved heptapeptide (YSPTSPS) sequences that are dynamically phosphorylated at the different stages of transcription [19]. Transient transfection of CTD-deleted pol II completely abrogated splicing of a β-globin reporter [18], suggesting an important role for the CTD in co-transcriptional splicing. This notion was further strengthened with the development of fluorescent microscopy. Misteli et al. used live cell imaging to demonstrate the relocalization of a fluorescently labeled splicing factor from nuclear speckles to sites of new transcription initiated from a β-tropomyosin minigene [20]. Furthermore, they and others have showed that such splicing factor mobilization is dependent on transcription through RNA polymerase II with an intact CTD [21–24]. More recently, several kinetic studies have shown that the vast majority of splicing in higher eukaryotes occurs co-transcriptionally, in a general 5′–3′ order. Altogether, these studies provide a rationale framework for coupling between elements at the transcribed DNA template and the splicing machinery.
2.2. Kinetic Regulation of Splicing
Almost immediately preceding the formal demonstration of co-transcriptional splicing, Kornblihtt and others revealed an intriguing new connection between these two processes that would ultimately pave the way for a new era in the alternative splicing field. In what would come to be known as the “kinetic” hypothesis, the rate of pol II elongation was implicated in alternative splicing decisions. As a first hint of things to come, it was demonstrated that promoter identity influences splicing decisions. Swapping promoters in minigene constructs resulted in altered splicing of weak exons [25], and recruitment of the splicing factor SF2/ASF (now SRSF1) to enhancers was shown to be promoter dependent [26]. These promoter-related splicing effects were ultimately attributed to elongation rate [27]. In support of a kinetic connection, the Smith group further found that inserting a binding site for the zinc-finger protein MAZ downstream of a weak exon led to pol II pausing and increased inclusion of the exon in spliced mRNA [28]. It was thus proposed that alternative splicing decisions are influenced by a temporal window of opportunity. Subsequent work fully established this notion and ultimately revealed a connection to chromatin (discussed below). Focusing on the fibronectin gene, inhibiting pol II elongation rate through use of a chemical inhibitor, 5,6-dichloro-1-beta-D ribofuranosylbenzimidazole (DRB), increased inclusion of the weak EDI exon, whereas increasing elongation rate through treatment with the histone deacetylase inhibitor trichostatin A (TSA) decreased exon inclusion [71]. A similar increase in EDI inclusion was seen in response to transcription with α-amanitin-resistant pol II with reduced elongation rate [10]. The global relevance of kinetic coupling between transcription and pre-mRNA splicing was later solidified through several genome-wide studies. Consistent throughout these studies, a slow elongation rate was associated with weak exon inclusion. For example, inhibition of pol II elongation with DRB or camptothecin in activated human T cells favored inclusion of weak exons in a subset of genes that were enriched in RNA processing and apoptosis pathways [29]. Similarly, hyperphosphorylation of the pol II CTD by ultraviolet irradiation (UV) and consequent inhibition of elongation rate induced alternative splicing of many genes involved in the DNA damage response [30].
Various models have been put forth to address the physiological signals for variable pol II elongation. For example, polymerase itself may be altered due to differential CTD phosphorylation or interaction with accessory factors [31] (Chapter 6). Alternatively, recent evidence suggests that the chromatin template of transcribed genes may provide polymerase with signals that locally regulate elongation rate, or provide direct information to the spliceosome regarding the location of exons. We expand on both these modes of chromatin-regulated splicing below.
2.3. Chromatin-Mediated Regulation of Splicing
Building on the accumulating evidence in support of co-transcriptional mRNA processing [32, 33], the first hints of a role for chromatin in splicing decisions date back to the early 1990s. Beckmann and Trifonov unexpectedly discovered that the average distance between the 3′ and 5′ splice sites flanking an exon followed a periodic pattern that was very close to the length covered by a single nucleosome [34]. Based on these results, they rationalized that placement of nucleosomes according to exon–intron boundaries may reflect an unanticipated role for chromatin structure in pre-mRNA splicing [34]. At the same time, an involvement for chromatin was also suspected when integrated copies of the adenovirus genome at distinct genomic locations in the same nuclei yielded different splicing outcomes [35]. Given that the viral genomes and cellular contexts were identical, the authors speculated that the chromatin structure at the integration site was responsible for the distinct splicing patterns, possibly through influencing the pol II elongation rate [35]. While these ideas gained momentum in the following years, it wasn’t until the advent of genome-wide sequencing data showing distinct chromatin patterns at exonic relative to intronic sequence that chromatin was solidified as a genuine contributor to splicing regulation. Genome-wide studies revealed that exons show a higher rate of nucleosome occupancy, specific histone modifications, and elevated DNA methylation relative to introns, raising the possibility that chromatin may aid the spliceosome in the process of exon definition. We discuss each of these associations in turn below.
2.3.1. Nucleosomes
A fundamental aspect of gene regulation in eukaryotes is the packaging of DNA into higher-order structures. In addition to maintaining genome integrity, this allows for the control of gene expression through the adoption of either transcriptionally inaccessible heterochromatin or relatively open euchromatin [36]. The basic building block of chromatin is the nucleosome, which is composed of approximately 146 base pairs of DNA wound around an octamer of histone proteins [37]. Canonical nucleosomes include two copies each of H2A, H2B, H3, and H4 histones [37]. Nucleosomes are inherent barriers to pol II elongation, as has been effectively demonstrated in vitro [38–40]. In order to accomplish efficient elongation in vivo, a variety of proteins cooperate to disassemble nucleosomes in front of elongating pol II and reestablish them in its wake [11, 41, 42]. Nucleosome turnover is a critical aspect of transcription fidelity, as nucleosome depletion facilitates unwanted cryptic transcription [43–47]. Given these very basic roles for nucleosomes in transcription, it was surprising to find a nonrandom intragenic positioning pattern across the genome: nucleosome occupancy is elevated at exons relative to introns, irrespective of gene expression status [48–50]. Considering the in vitro data demonstrating that nucleosomes are barriers to pol II elongation, these observations suggested that nucleosomes promote exon definition by facilitating transient pol II pausing and consequent spliceosome assembly. While not formally demonstrated due to difficulties in depleting nucleosomes without untoward effects on gene expression, several studies support this premise. For example, exons with weak splice sites show higher nucleosome density compared to constitutive exons, and pseudoexons with strong splice sites are nucleosome depleted [48]. Furthermore, the average size of a mammalian exon (145 base pairs) is similar to the length of DNA wrapped within a single nucleosome [48, 51]. Additionally, a recent study suggested a role for variant histone incorporation in splicing regulation. Depletion of the mammalian-specific H2A variant, H2A.Bbd, which is preferentially found in transcribed sequences, led to a global decrease in splicing efficiency [52]. As several proteins involved in RNA processing as well as major splicing components were coprecipitated with H2A.Bbd [52–55], these studies hint at a direct role for nucleosomes in splicing regulation. This notion is further reinforced by evidence implicating posttranslational modification of histone tails in splicing regulation, as described below.
An additional twist in the interplay between nucleosomes and splicing is related to nucleotide content. Exonic sequences tend to be GC-rich as compared to introns, which inherently favors nucleosome deposition at exons [56, 57]. Nucleotide bias has been implicated in alternative splicing regulation [58], and a recent study identified a role for the DBIRD complex (composed of deleted in breast cancer 1 [DBC1] and ZNF236) in the exclusion of AT-rich exons. Through a yet undefined mechanism, DBIRD interacts with pol II and facilitates efficient passage through AT-rich sequence. In the absence of DBIRD, polymerase stalls at the weak exons and leads to increased inclusion [59]. This study highlights an additional emerging theme: splicing, chromatin, and transcription are highly intertwined.
2.3.2. Modification of Histone Tails
The advent of high-throughput deep sequencing following chromatin immunoprecipitation (ChIP-seq) has allowed for the dissection of the intragenic epigenome across diverse species, including human, mice, C. elegans, and Drosophila [50, 51, 60, 61]. Several histone marks were found to be particularly prevalent on exonic sequences, including trimethylation of H3 lysine 36 (H3K36me3) [47, 62–65], dimethylation of H3 lysine 27 (H3K27me2) [49, 64], and monomethylation of H3 lysine 79 (H3K79me1) and H2B lysine 5 (H2BK5me1) [62, 63]. In contrast, H3K9 methylation is depleted at exons [66]. Intriguingly, exonic enrichment is not evenly distributed across gene bodies. For example, H3K36me3 levels rise into gene bodies, whereas H3K4me3 is found near transcription start sites [60]. H3K36me3, in particular, has received significant attention as it is exclusively found at actively transcribed genes, suggesting an important role in pre-mRNA processing [60]. Acknowledging that histone modification measurements must be adjusted for the overall increase in nucleosome content at exonic sequences, H3K36me3 enrichment at exons persists after nucleosome correction [64]. While the genome-wide associations have largely focused on histone methylation, additional modifications to histone tails have the potential to influence splicing. For example, a recent study in yeast reported elevated monoubiquitylation of H2B lysine 123 (H2BK123ub1) in introns of transcribed genes, and disruption of H2BK123ub1 altered the distribution of H3K6me3 [67], suggesting functional antagonism in exon–intron definition through specific posttranslational modifications of histone tails.
A direct role for histone modifications in exon definition is supported by a variety of studies involving modulation of specific histone posttranslational marks. As a result, two non-exclusive potential mechanisms by which chromatin can influence alternative splicing decisions have been proposed: (1) local alteration of pol II elongation rate and (2) site-specific recruitment of RNA-binding proteins through interaction with chromatin-binding proteins. Examples of both modes of regulation exist and are largely intertwined as described below.
Histone Modifications in Kinetic Regulation
The strongest evidence in favor of a role for chromatin in kinetic regulation of splicing comes from studies in which exogenous stimuli led to global or local changes in chromatin structure and associated changes in splicing. In general, acetylation of histone tails is associated with an open chromatin context and efficient pol II processivity [68]. Several studies from the Kornblihtt group have shown that tipping the “accessibility” balance in either direction alters exon inclusion of both model genes and genome wide. For example, membrane depolarization of neuronal cells led to increased H3K9 acetylation (H3K9ac) specifically within the vicinity of exon 18 of the NCAM gene and promoted exclusion of the exon from spliced mRNA [69]. This exon appears to be particularly susceptible to kinetic regulation, as evidenced through use of a mutant polymerase with reduced elongation rate. In addition, globally increasing acetylation with the histone deacetylase inhibitor trichostatin A (TSA) decreased exon 18 inclusion [70]. A similar decrease in the fibronectin EDI exon was seen following TSA treatment [71]. In contrast, the opposing, repressive modification, H3K9 methylation, has been shown to mediate exon inclusion in a number of systems. The Muchardt group defined a role for H3K9me3 in alternative splicing of CD44 pre-mRNA and was further able to uncover a potential physiological link to polymerase pausing. In response to phorbol 12-myristate 13-acetate (PMA) treatment, CD44 transcripts show increased inclusion of nine tandem alternative exons, which is associated with increased pol II occupancy and a local increase in H3K9me3 detection [72, 73]. Stimulation also resulted in increased detection of the H3K9me3-interacting chromodomain protein HP1γ at the alternative exons. Remarkably, RNAi-mediated depletion of HP1γ abrogated both pol II accumulation and variant exon inclusion [74]. This study suggested that H3K9me3-associated HP1γ somehow bridged the processes of transcription and splicing. Indeed, it has been reported that HP1γ is enriched at hundreds of active genes and promotes co-transcriptional splicing through recruitment of the spliceosomal protein U1–70K and SR family protein SRSF1 [75, 76].
Additional evidence supporting a role for H3K9 methylation in the kinetic regulation of splicing comes from studies involving exogenous introduction of siRNAs directed against intragenic sequence. Analogous to Argonaute protein-dependent transcriptional gene silencing (TGS), wherein siRNA directed against promoter DNA triggers gene silencing through local heterochromatin formation [77–80], extension of TGS into intragenic sequence mediates chromatin changes that locally modulate pol II elongation without affecting overall gene expression. For example, exogenous siRNAs targeted against an intronic sequence proximal to the fibronectin EDI exon in human cells promoted an AGO1-dependent local increase in the repressive chromatin marks H3K9me2 and H3K27me3 and increased exon inclusion [81]. The authors further demonstrated that the altered chromatin structure resulted in HP1α recruitment to the region, suggesting a similar bridging effect as described for HP1γ above. A recent study in the CD44 model system also directly implicated the Argonaute proteins AGO1 and AGO2 in bridging interactions. AGO1 and AGO2 interact with splicing factors and are recruited to the CD44 variant exons in a Dicer and HP1γ dependent–manner, culminating in increased exon inclusion through reduced pol II processivity [82]. These complex associations were reverberated in several studies from the Kennedy laboratory examining the Argonaute-related nuclear RNAi defective (NRDE) protein-dependent intragenic TGS pathway in C. elegans. Both exogenous and endogenous siRNA-associated recruitment of NRDE factors were found to promote accumulation of H3K9me3 and inhibited pol II elongation in C. elegans [83, 84]. Furthermore, endogenous siRNAs directed NRDE-1 to interact with both chromatin and pre-mRNA, thereby revealing a conserved role for Argonaute proteins in connecting these nuclear processes [84, 85].
As is evident throughout these studies, post-transcriptional modification of H3K9 seems to have a central role in the kinetic regulation of splicing. It is worth noting that these observations are somewhat at odds with the intergenic role of H3K9. Trimethylation of H3K9 is a classic feature of heterochromatin formation and is associated with repeat elements and otherwise silenced areas of the genome [86]. However, intragenic H3K9me3 is not strictly associated with transcriptional repression [87]. These studies suggest that H3K9 may mediate context-dependent effects on transcription and/or splicing. It will certainly be interesting to examine the role of this modification in splicing in greater detail in the coming years.
Histone Modifications in Adaptor Function
While several aspects of bridging from chromatin to RNA were highlighted in the discussion of kinetic regulation above, it is unlikely that all alternative exons are strictly under kinetic regulation. Indeed, studies focused on the role of H3K36 and H3K4 methylation in splicing reveal a more general role for chromatin modifications as adaptors for RNA-binding protein recruitment. Interestingly, as shown for H3K36me3, the same chromatin modification can be associated with distinct splicing outcomes dependent on the interacting factors. For example, the Misteli laboratory has shown that H3K36me3 is associated with exclusion of a subset of PTB-dependent exons. PTB is recruited to these exons through interaction with the H3K36me3-binding protein, MRG15 [88]. As a proof of principle, depletion of H3K36me3 levels through RNAi against the K36 methyltransferase, Setd2, increased the inclusion level of these exons [88]. In contrast, the Bickmore laboratory has shown that H3K36me3 is associated with inclusion of a subset of exons due to Psip1-/Ledgf-dependent recruitment of the splicing factor SRSF1. In a strikingly similar mechanism, Psip1 interacts with H3K36me3 and recruits the positive acting splicing factor to a subset of exons, and SRSF1 localization and splicing are altered in Psip1 null cells [89]. These contrasting studies illustrate the clear involvement of additional context-dependent factors that remain to be identified.
Adaptor function has also been demonstrated for H3K4me3, which is enriched at the 5′ ends of active genes. Biochemical purification identified CHD1 as an H3K4me3-interacting protein and CHD1 was also found to interact with the spliceosomal proteins U2 snRNP [90]. Depletion of either CHD1 or H3K4me3 through RNAi reduced U2 association with chromatin and reduced pre-mRNA splicing efficiency [90]. Similarly, the histone 3 acetyltransferase, GCN5, promotes co-transcriptional U2 snRNP recruitment [91], hinting at a possible adaptor function. Altogether, the sum of these studies demonstrates a clear role for chromatin structure in constitutive and alternative splicing regulation, both through kinetic and adaptor mechanisms.
2.4. DNA Methylation
As introduced above, DNA methylation also shows a nonrandom intragenic distribution pattern: methylation levels are significantly enhanced at exonic relative to intronic sequences. However, unlike promoter DNA methylation, which is associated with gene silencing [92], methylation within gene bodies does not have a clear relationship to gene expression levels [93]. Instead, genome-wide methylome analyses in lower eukaryotes foreshadowed a potential role for DNA methylation in splicing regulation. Comparative methylome analyses from the Jacobsen and Zilberman laboratories showed that the acquisition of DNA methylation predates the divergence of plant and animal lineages, and revealed conserved enrichment at exonic sequences relative to introns [94, 95]. In comparing genetically identical honeybee castes, the Maleszka laboratory further showed that differences in queen versus worker bee methylome patterns correlate with changes in alternative splicing patterns. Strikingly, they also found that the low level of DNA methylation is seemingly restricted to exons and is depleted at intronless genes [96]. Subsequent work in additional insect model systems has confirmed an association between intragenic methylation, alternative splicing, and phenotypic diversity [97–99]. While mammalian methylomes are comparatively complex and widespread intergenic methylation is found, these intragenic features are highly conserved [100, 101]. High-resolution bisulfite sequencing of the human genome validated the enrichment of exonic methylation and revealed sharp transitions at exon–intron junctions [102]. A reanalysis of several genome-wide human methylome and RNA-seq datasets established that methylation levels correlate with alternative splicing in human cells. Included exons showed an overall higher level of DNA methylation relative to excluded exons, suggesting a direct role for DNA methylation in exon definition. These associations persisted even after correcting for increased nucleosome and GC content at exons relative to introns [93].
The conserved association between exonic DNA methylation and alternative splicing across diverse taxa suggests regulated mechanisms for the establishment and removal of methylation patterns. While the mechanisms by which DNA methyltransferases (DNMTs) are targeted to exons at distinct stages in development remain unknown, recent studies have begun to reveal a basis for variable DNA methylation. 5-methylcytosine (5-mC) can be converted to 5-hydroxymethylcytosine (5-hmC) through the activity of the TET family of proteins. 5-hmC can stably persist in the genome, or can be further converted to unmethylated cytosine through additional oxidation. Notably, bisulfite sequencing is unable to distinguish 5-mC from 5-hmC, and recent studies aimed at specifically deciphering the 5-hmC methylome have found an overlapping distribution pattern: 5-hmC is also enriched at exons relative to introns [103, 104]. Furthermore, 5-hmC levels were also shown to undergo sharp transitions at exon–intron boundaries in the brain, and alternative exons showed an overall lower level of 5-hmC relative to constitutive exons. Curiously, non-neural tissues showed more 5-mC at exon–intron boundaries [105]. These studies suggest that tissue-specific changes in the ratio of 5-mC to 5-hmC may represent a novel mode of alternative splicing regulation.
While the accumulation of these genome-wide data over the last several years strongly suggested a fundamental role for DNA methylation in exon definition, potential mechanisms had remained more elusive. Given that the majority of splicing occurs co-transcriptionally, possibilities included direct impact of DNA methylation on splicing through kinetic regulation or indirect regulation through variable interaction with auxiliary factors. We recently provided evidence for the latter possibility. Through our analysis of alternative splicing of CD45 pre-mRNA, we determined that inclusion of variable exon 5 is mediated by reciprocal binding of the zinc-finger protein, CTCF, and 5-methylcytosine. The binding of CTCF to CD45 DNA acts as a transient barrier to pol II elongation, which kinetically favors spliceosome assembly at the weak splice sites. In contrast, DNA methylation acts to evict CTCF and thereby abolishes pol II pausing and exon 5 inclusion. CTCF-ChIP-seq and RNA-seq in CTCF-depleted cells verified CTCF to be a global regulator of alternative splicing [106]. This study provided the first mechanistic link between DNA methylation and alternative splicing. A similar effect was recently shown for the zinc-finger protein VEZF1: binding of VEZF1 to DNA promotes pol II pausing and results in alternative splicing of a subset of genes [107]. Interestingly, like CTCF, VEZF1 interaction with DNA protects against DNA methylation [108]. In addition, VEZF1 interacts with MRG15, which was previously implicated to be a chromatin-binding adaptor between H3K36me3 and PTB as described above [88, 107]. Together, these studies suggest a basic role for intragenic binding of zinc-finger proteins in kinetically regulated splicing. We predict that many additional examples of DNA-binding regulators of splicing will be revealed in the coming years.
It should further be noted that we found CD45 exon 5 methylation to be developmentally regulated. Naïve peripheral lymphocytes show enhanced CTCF binding, whereas mature lymphocytes show enhanced exon 5 methylation. Thus, mechanisms certainly exist that promote exon 5 methylation in a stage-specific manner. Support for active remodeling of intragenic methylation can also be found in the honeybee genome. As mentioned above, genetically identical queen and worker honeybee show distinct methylation profiles. Remarkably, RNAi-mediated depletion of the de novo methyltransferase DNMT3 generated adult bees with queen-like characteristics, and queen-specific isoforms have been defined that are distinguished by unique overlapping exonic DNA methylation [96, 109]. Given that DNA methylation is not strictly associated with exon inclusion or exclusion [110], it is possible that intragenic DNA methylation plays a fundamental role in developmentally regulated alternative splicing through association with a complex network of methyl-resistant and methyl-sensitive DNA-binding proteins.
3. Splicing Reciprocally Modulates Chromatin Structure
While the role of chromatin structure in splicing is now well established, an emerging area of coupling is recent evidence showing that splicing can reciprocally influence chromatin modifications. Independent publications from the Carmo-Fonseca and Bentley laboratories indicated that H3K36me3 deposition over gene bodies is splicing dependent [111, 112]. The Bentley laboratory showed that mutation of 3′ splice sites in the upstream introns of a β-globin reporter resulted in repositioning of H3K36me3 from the 5′ to the 3′ end of the reporter. Globally inhibiting splicing with spliceostatin A also resulted in H3K36me3 redistribution further 3′ into genes [111]. The Carmo-Fonseca group further showed that splicing promotes recruitment of the H3K36 methyltransferase HYPB/Setd2 to gene bodies. Inhibition of splicing globally reduced H3K36me3, whereas activating splicing of a model gene had the opposite effect. Furthermore, intronless genes show lower H3K6me3 levels irrespective of expression status [112], implicating splicing rather than general transcription in Setd2 recruitment. The notion that pre-mRNA processing can influence histone modification is further strengthened by the demonstration that the RNA-binding Hu proteins can modulate histone acetylation. The binding of Hu proteins to target sites on mRNA flanking alternative exons led to local inhibition of histone deacetylase 2 (HDAC2) activity and increased histone acetylation and exon exclusion [113]. Altogether, these studies suggest that spliceosomal components actively connect histone-modifying enzymes to transcription through yet unknown mechanisms.
4. Concluding Remarks
As highlighted throughout this chapter, the last several years have revealed an extensive network of coupling between pre-mRNA splicing and chromatin. Studies of model genes and genome-wide analyses have revealed exonic epigenetic signatures. Some of these signatures, such as nucleosomes [48–50], are found independent of transcription status, whereas others, such as H3K36me3, are transcription and splicing dependent [111, 112]. Chromatin modifications have been shown to influence spliceosome assembly through kinetic regulation of pol II elongation and through recruitment of splicing factors to their required sites of action [11]. An area that will be particularly interesting to follow in the coming years is how these changes in chromatin structure are modulated during development. While the current studies have focused on the chromatin template, pol II, and the spliceosome, in reality, a vast network of remodelers is required to effect chromatin changes. For example, how are constitutively expressed histone- and DNA-modifying enzymes targeted to intragenic sequences at specific stages in development? In addition, do additional chromatin-associated factors that are critical for transcription, such as histone chaperones, have a role in splicing regulation? Certainly, our current understanding of chromatin-mediated mRNA splicing field is hazy at best, and future research in this area is likely to be full of surprises.
References
- 1.Black DL (2003) Mechanisms of alternative pre-messenger RNA splicing. Annu Rev Biochem 72:291–336. doi: 10.1146/annurev.biochem.72.121801.161720, 121801.161720[pii] [DOI] [PubMed] [Google Scholar]
- 2.Wang ET, Sandberg R, Luo S et al. (2008) Alternative isoform regulation in human tissue transcriptomes. Nature 456(7221): 470–476. doi: nature07509, [pii] 10.1038/nature07509 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Pan Q, Shai O, Lee LJ et al. (2008) Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat Genet 40(12):1413–1415. doi: ng.259, [pii] 10.1038/ng.259 [DOI] [PubMed] [Google Scholar]
- 4.Busch A, Hertel KJ (2011) Evolution of SR protein and hnRNP splicing regulatory factors. Wiley Interdiscip Rev RNA. doi: 10.1002/wrna.100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Matlin AJ, Clark F, Smith CW (2005) Understanding alternative splicing: towards a cellular code. Nat Rev Mol Cell Biol 6(5): 386–398 [DOI] [PubMed] [Google Scholar]
- 6.Allo M, Schor IE, Munoz MJ et al. (2010) Chromatin and alternative splicing. Cold Spring Harb Symp Quant Biol 75:103–111. doi: sqb.2010.75.023, [pii] 10.1101/sqb.2010.75.023 [DOI] [PubMed] [Google Scholar]
- 7.Oesterreich FC, Bieberstein N, Neugebauer KM (2011) Pause locally, splice globally. Trends Cell Biol 21(6):328–335. doi: S0962–8924(11)00035–3, [pii] 10.1016/j.tcb.2011.03.002 [DOI] [PubMed] [Google Scholar]
- 8.Han J, Xiong J, Wang D et al. (2011) Pre-mRNA splicing: where and when in the nucleus. Trends Cell Biol 21(6):336–343. doi: S0962–8924(11)00036–5, [pii] 10.1016/j.tcb.2011.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Egloff S, Murphy S (2008) Cracking the RNA polymerase II CTD code. Trends Genet 24(6):280–288. doi: S0168–9525(08)00128–5, [pii] 10.1016/j.tig.2008.03.008 [DOI] [PubMed] [Google Scholar]
- 10.de la Mata M, Alonso CR, Kadener S et al. (2003) A slow RNA polymerase II affects alternative splicing in vivo. Mol Cell 12(2): 525–532. doi: S1097276503003101 [pii] [DOI] [PubMed] [Google Scholar]
- 11.Shukla S, Oberdoerffer S (2012) Co-transcriptional regulation of alternative pre-mRNA splicing. Biochim Biophys Acta 1819(7):673–683. doi: S1874–9399(12)00032–6, [pii] 10.1016/j.bbagrm.2012.01.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Beyer AL, Bouton AH, Miller OL Jr (1981) Correlation of hnRNP structure and nascent transcript cleavage. Cell 26(2 Pt 2):155–165 [DOI] [PubMed] [Google Scholar]
- 13.Beyer AL, Osheim YN (1988) Splice site selection, rate of splicing, and alternative splicing on nascent transcripts. Genes Dev 2(6):754–765 [DOI] [PubMed] [Google Scholar]
- 14.Dower K, Rosbash M (2002) T7 RNA polymerase-directed transcripts are processed in yeast and link 3′ end formation to mRNA nuclear export. RNA 8(5):686–697 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.McCracken S, Rosonina E, Fong N et al. (1998) Role of RNA polymerase II carboxy-terminal domain in coordinating transcription with RNA processing. Cold Spring Harb Symp Quant Biol 63:301–309 [DOI] [PubMed] [Google Scholar]
- 16.Sisodia SS, Sollner-Webb B, Cleveland DW (1987) Specificity of RNA maturation pathways: RNAs transcribed by RNA polymerase III are not substrates for splicing or polyadenylation. Mol Cell Biol 7(10):3602–3612 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Smale ST, Tjian R (1985) Transcription of herpes simplex virus tk sequences under the control of wild-type and mutant human RNA polymerase I promoters. Mol Cell Biol 5(2): 352–362 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.McCracken S, Fong N, Yankulov K et al. (1997) The C-terminal domain of RNA polymerase II couples mRNA processing to transcription. Nature 385(6614):357–361. doi: 10.1038/385357a0 [DOI] [PubMed] [Google Scholar]
- 19.Dahmus ME (1996) Reversible phosphorylation of the C-terminal domain of RNA polymerase II. J Biol Chem 271(32):19009–19012 [DOI] [PubMed] [Google Scholar]
- 20.Misteli T, Caceres JF, Spector DL (1997) The dynamics of a pre-mRNA splicing factor in living cells. Nature 387(6632):523–527. doi: 10.1038/387523a0 [DOI] [PubMed] [Google Scholar]
- 21.de la Mata M, Alonso CR, Kadener S et al. (2003) A slow RNA polymerase II affects alternative splicing in vivo. Mol Cell 12(2): 525–532 [DOI] [PubMed] [Google Scholar]
- 22.de la Mata M, Kornblihtt AR (2006) RNA polymerase II C-terminal domain mediates regulation of alternative splicing by SRp20. Nat Struct Mol Biol 13(11):973–980. doi: 10.1038/nsmb1155 [DOI] [PubMed] [Google Scholar]
- 23.Listerman I, Sapra AK, Neugebauer KM (2006) Cotranscriptional coupling of splicing factor recruitment and precursor messenger RNA splicing in mammalian cells. Nat Struct Mol Biol 13(9):815–822. doi: 10.1038/nsmb1135 [DOI] [PubMed] [Google Scholar]
- 24.Misteli T, Spector DL (1999) RNA polymerase II targets pre-mRNA splicing factors to transcription sites in vivo. Mol Cell 3(6): 697–705 [DOI] [PubMed] [Google Scholar]
- 25.Cramer P, Pesce CG, Baralle FE et al. (1997) Functional association between promoter structure and transcript alternative splicing. Proc Natl Acad Sci USA 94(21):11456–11460 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Cramer P, Caceres JF, Cazalla D et al. (1999) Coupling of transcription with alternative splicing: RNA pol II promoters modulate SF2/ASF and 9G8 effects on an exonic splicing enhancer. Mol Cell 4(2):251–258. doi: S1097-2765(00)80372-X [pii] [DOI] [PubMed] [Google Scholar]
- 27.Kadener S, Fededa JP, Rosbash M et al. (2002) Regulation of alternative splicing by a transcriptional enhancer through RNA pol II elongation. Proc Natl Acad Sci USA 99(12):8185–8190. doi: 10.1073/pnas.122246099, 99/12/8185 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Roberts GC, Gooding C, Mak HY et al. (1998) Co-transcriptional commitment to alternative splice site selection. Nucleic Acids Res 26(24):5568–5572 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Ip JY, Schmidt D, Pan Q et al. (2011) Global impact of RNA polymerase II elongation inhibition on alternative splicing regulation. Genome Res 21(3):390–401. doi: 10.1101/gr.111070.110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Munoz MJ, Perez Santangelo MS, Paronetto MP et al. (2009) DNA damage regulates alternative splicing through inhibition of RNA polymerase II elongation. Cell 137(4):708–720. doi: 10.1016/j.cell.2009.03.010 [DOI] [PubMed] [Google Scholar]
- 31.Munoz MJ, de la Mata M, Kornblihtt AR (2010) The carboxy terminal domain of RNA polymerase II and alternative splicing. Trends Biochem Sci 35(9):497–504. doi: 10.1016/j.tibs.2010.03.010 [DOI] [PubMed] [Google Scholar]
- 32.Yue Z, Maldonado E, Pillutla R et al. (1997) Mammalian capping enzyme complements mutant Saccharomyces cerevisiae lacking mRNA guanylyltransferase and selectively binds the elongating form of RNA polymerase II. Proc Natl Acad Sci USA 94(24):12898–12903 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Hirose Y, Manley JL (1998) RNA polymerase II is an essential mRNA polyadenylation factor. Nature 395(6697):93–96. doi: 10.1038/25786 [DOI] [PubMed] [Google Scholar]
- 34.Beckmann JS, Trifonov EN (1991) Splice junctions follow a 205-base ladder. Proc Natl Acad Sci USA 88(6):2380–2383 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Adami G, Babiss LE (1991) DNA template effect on RNA splicing: two copies of the same gene in the same nucleus are processed differently. EMBO J 10(11):3457–3465 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Kouzarides T (2007) Chromatin modifications and their function. Cell 128(4):693–705. doi: S0092–8674(07)00184–5, [pii] 10.1016/j.cell.2007.02.005 [DOI] [PubMed] [Google Scholar]
- 37.Luger K, Mader AW, Richmond RK et al. (1997) Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389(6648):251–260. doi: 10.1038/38444 [DOI] [PubMed] [Google Scholar]
- 38.Knezetic JA, Luse DS (1986) The presence of nucleosomes on a DNA template prevents initiation by RNA polymerase II in vitro. Cell 45(1):95–104 [DOI] [PubMed] [Google Scholar]
- 39.Izban MG, Luse DS (1991) Transcription on nucleosomal templates by RNA polymerase II in vitro: inhibition of elongation with enhancement of sequence-specific pausing. Genes Dev 5(4):683–696 [DOI] [PubMed] [Google Scholar]
- 40.Bondarenko VA, Steele LM, Ujvari A et al. (2006) Nucleosomes can form a polar barrier to transcript elongation by RNA polymerase II. Mol Cell 24(3):469–479. doi: 10.1016/j.molcel.2006.09.009 [DOI] [PubMed] [Google Scholar]
- 41.Kristjuhan A, Svejstrup JQ (2004) Evidence for distinct mechanisms facilitating transcript elongation through chromatin in vivo. EMBO J 23(21):4243–4252. doi: 10.1038/sj.emboj.7600433, 7600433 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Schwabish MA, Struhl K (2004) Evidence for eviction and rapid deposition of histones upon transcriptional elongation by RNA polymerase II. Mol Cell Biol 24(23):10111–10117. doi: 24/23/10111, [pii] 10.1128/MCB.24.23.10111-10117.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Mason PB, Struhl K (2003) The FACT complex travels with elongating RNA polymerase II and is important for the fidelity of transcriptional initiation in vivo. Mol Cell Biol 23(22):8323–8333 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Schwabish MA, Struhl K (2006) Asf1 mediates histone eviction and deposition during elongation by RNA polymerase II. Mol Cell 22(3):415–422. doi: S1097–2765(06)00180–8, [pii] 10.1016/j.molcel.2006.03.014 [DOI] [PubMed] [Google Scholar]
- 45.Kaplan CD, Laprade L, Winston F (2003) Transcription elongation factors repress transcription initiation from cryptic sites. Science 301(5636):1096–1099. doi: 10.1126/science.1087374, 301/5636/1096 [pii] [DOI] [PubMed] [Google Scholar]
- 46.Carrozza MJ, Li B, Florens L, Suganuma T et al. (2005) Histone H3 methylation by Set2 directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic transcription. Cell 123(4):581–592. doi: S0092–8674(05)01156–6, [pii] 10.1016/j.cell.2005.10.023 [DOI] [PubMed] [Google Scholar]
- 47.Kolasinska-Zwierz P, Down T, Latorre I et al. (2009) Differential chromatin marking of introns and expressed exons by H3K36me3. Nat Genet 41(3):376–381. doi: ng.322, [pii] 10.1038/ng.322 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Tilgner H, Nikolaou C, Althammer S et al. (2009) Nucleosome positioning as a determinant of exon recognition. Nat Struct Mol Biol 16(9):996–1001. doi: 10.1038/nsmb.1658 [DOI] [PubMed] [Google Scholar]
- 49.Andersson R, Enroth S, Rada-Iglesias A et al. (2009) Nucleosomes are well positioned in exons and carry characteristic histone modifications. Genome Res 19(10):1732–1741. doi: 10.1101/gr.092353.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Spies N, Nielsen CB, Padgett RA et al. (2009) Biased chromatin signatures around polyadenylation sites and exons. Mol Cell 36(2):245–254. doi: 10.1016/j.molcel.2009.10.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Schwartz S, Meshorer E, Ast G (2009) Chromatin organization marks exon-intron structure. Nat Struct Mol Biol 16(9):990–995. doi: 10.1038/nsmb.1659 [DOI] [PubMed] [Google Scholar]
- 52.Tolstorukov MY, Goldman JA, Gilbert C et al. (2012) Histone variant H2A.Bbd is associated with active transcription and mRNA processing in human cells. Mol Cell 47(4):596–607. doi: 10.1016/j.molcel.2012.06.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Corrionero A, Minana B, Valcarcel J (2011) Reduced fidelity of branch point recognition and alternative splicing induced by the antitumor drug spliceostatin A. Genes Dev 25(5):445–459. doi: 10.1101/gad.2014311 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Folco EG, Coil KE, Reed R (2011) The antitumor drug E7107 reveals an essential role for SF3b in remodeling U2 snRNP to expose the branch point-binding region. Genes Dev 25(5):440–444. doi: 10.1101/gad.2009411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Gozani O, Feld R, Reed R (1996) Evidence that sequence-independent binding of highly conserved U2 snRNP proteins upstream of the branch site is required for assembly of spliceosomal complex A. Genes Dev 10(2):233–243 [DOI] [PubMed] [Google Scholar]
- 56.Bernardi G (1993) The vertebrate genome: isochores and evolution. Mol Biol Evol 10(1):186–204 [DOI] [PubMed] [Google Scholar]
- 57.Peckham HE, Thurman RE, Fu Y et al. (2007) Nucleosome positioning signals in genomic DNA. Genome Res 17(8):1170–1177. doi: 10.1101/gr.6101007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Amit M, Donyo M, Hollander D et al. (2012) Differential GC content between exons and introns establishes distinct strategies of splice-site recognition. Cell Rep 1(5):543–556. doi: 10.1016/j.celrep.2012.03.013 [DOI] [PubMed] [Google Scholar]
- 59.Close P, East P, Dirac-Svejstrup AB et al. (2012) DBIRD complex integrates alternative mRNA splicing with RNA polymerase II transcript elongation. Nature 484(7394):386–389. doi: 10.1038/nature10925 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Kolasinska-Zwierz P, Down T, Latorre I et al. (2009) Differential chromatin marking of introns and expressed exons by H3K36me3. Nat Genet 41(3):376–381. doi: 10.1038/ng.322 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Huff JT, Plocik AM, Guthrie C et al. (2010) Reciprocal intronic and exonic histone modification regions in humans. Nat Struct Mol Biol 17(12):1495–1499. doi: 10.1038/nsmb.1924 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Andersson R, Enroth S, Rada-Iglesias A et al. (2009) Nucleosomes are well positioned in exons and carry characteristic histone modifications. Genome Res 19(10):1732–1741. doi: gr.092353.109, [pii] 10.1101/gr.092353.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Schwartz S, Meshorer E, Ast G (2009) Chromatin organization marks exon-intron structure. Nat Struct Mol Biol 16(9): 990–995. doi: nsmb.1659, [pii] 10.1038/nsmb.1659 [DOI] [PubMed] [Google Scholar]
- 64.Spies N, Nielsen CB, Padgett RA et al. (2009) Biased chromatin signatures around polyadenylation sites and exons. Mol Cell 36(2):245–254. doi: S1097–2765(09)00743–6, [pii] 10.1016/j.molcel.2009.10.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Huff JT, Plocik AM, Guthrie C et al. (2010) Reciprocal intronic and exonic histone modification regions in humans. Nat Struct Mol Biol 17(12):1495–1499. doi: nsmb.1924, [pii] 10.1038/nsmb.1924 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Dhami P, Saffrey P, Bruce AW et al. (2010) Complex exon-intron marking by histone modifications is not determined solely by nucleosome distribution. PLoS One 5(8):e12339. doi: 10.1371/journal.pone.0012339 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Shieh GS, Pan CH, Wu JH et al. (2011) H2B ubiquitylation is part of chromatin architecture that marks exon-intron structure in budding yeast. BMC Genomics 12:627. doi: 10.1186/1471-2164-12-627 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Hnilicova J, Hozeifi S, Duskova E et al. (2011) Histone deacetylase activity modulates alternative splicing. PLoS One 6(2):e16727. doi: 10.1371/journal.pone.0016727 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Schor IE, Rascovan N, Pelisch F et al. (2009) Neuronal cell depolarization induces intragenic chromatin modifications affecting NCAM alternative splicing. Proc Natl Acad Sci USA 106(11):4325–4330. doi: 10.1073/pnas.0810666106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Schor IE, Rascovan N, Pelisch F et al. (2009) Neuronal cell depolarization induces intragenic chromatin modifications affecting NCAM alternative splicing. Proc Natl Acad Sci USA 106(11):4325–4330. doi: 0810666106, [pii] 10.1073/pnas.0810666106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Nogues G, Kadener S, Cramer P et al. (2002) Transcriptional activators differ in their abilities to control alternative splicing. J Biol Chem 277(45):43110–43114. doi: 10.1074/jbc.M208418200 [DOI] [PubMed] [Google Scholar]
- 72.Batsche E, Yaniv M, Muchardt C (2006) The human SWI/SNF subunit Brm is a regulator of alternative splicing. Nat Struct Mol Biol 13(1):22–29. doi: nsmb1030, [pii] 10.1038/nsmb1030 [DOI] [PubMed] [Google Scholar]
- 73.Saint-Andre V, Batsche E, Rachez C et al. (2011) Histone H3 lysine 9 trimethylation and HP1gamma favor inclusion of alternative exons. Nat Struct Mol Biol 18(3):337–344. doi: nsmb.1995, [pii] 10.1038/nsmb.1995 [DOI] [PubMed] [Google Scholar]
- 74.Saint-Andre V, Batsche E, Rachez C et al. (2011) Histone H3 lysine 9 trimethylation and HP1gamma favor inclusion of alternative exons. Nat Struct Mol Biol 18(3):337–344. doi: 10.1038/nsmb.1995 [DOI] [PubMed] [Google Scholar]
- 75.Smallwood A, Hon GC, Jin F et al. (2012) CBX3 regulates efficient RNA processing genome-wide. Genome Res 22(8):1426–1436. doi: 10.1101/gr.124818.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Greil F, van der Kraan I, Delrow J et al. (2003) Distinct HP1 and Su(var)3–9 complexes bind to sets of developmentally coexpressed genes depending on chromosomal location. Genes Dev 17(22):2825–2838. doi: 10.1101/gad.281503 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Janowski BA, Huffman KE, Schwartz JC et al. (2006) Involvement of AGO1 and AGO2 in mammalian transcriptional silencing. Nat Struct Mol Biol 13(9):787–792. doi: 10.1038/nsmb1140 [DOI] [PubMed] [Google Scholar]
- 78.Morris KV, Chan SW, Jacobsen SE et al. (2004) Small interfering RNA-induced transcriptional gene silencing in human cells. Science 305(5688):1289–1292. doi: 10.1126/science.1101372 [DOI] [PubMed] [Google Scholar]
- 79.Kim DH, Villeneuve LM, Morris KV et al. (2006) Argonaute-1 directs siRNA-mediated transcriptional gene silencing in human cells. Nat Struct Mol Biol 13(9):793–797. doi: 10.1038/nsmb1142 [DOI] [PubMed] [Google Scholar]
- 80.Weinberg MS, Villeneuve LM, Ehsani A et al. (2006) The antisense strand of small interfering RNAs directs histone methylation and transcriptional gene silencing in human cells. RNA 12(2):256–262. doi: 10.1261/rna.2235106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Allo M, Buggiano V, Fededa JP et al. (2009) Control of alternative splicing through siRNA-mediated transcriptional gene silencing. Nat Struct Mol Biol 16(7):717–724. doi: 10.1038/nsmb.1620 [DOI] [PubMed] [Google Scholar]
- 82.Ameyar-Zazoua M, Rachez C, Souidi M et al. (2012) Argonaute proteins couple chromatin silencing to alternative splicing. Nat Struct Mol Biol. doi: 10.1038/nsmb.2373 [DOI] [PubMed] [Google Scholar]
- 83.Guang S, Bochner AF, Burkhart KB et al. (2010) Small regulatory RNAs inhibit RNA polymerase II during the elongation phase of transcription. Nature 465(7301):1097–1101. doi: 10.1038/nature09095 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Burkhart KB, Guang S, Buckley BA et al. (2011) A pre-mRNA-associating factor links endogenous siRNAs to chromatin regulation. PLoS Genet 7(8):e1002249. doi: 10.1371/journal.pgen.1002249, PGENETICS-D-11–00476 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Ameyar-Zazoua M, Rachez C, Souidi M et al. (2012) Argonaute proteins couple chromatin silencing to alternative splicing. Nat Struct Mol Biol 19(10):998–1004. doi: nsmb.2373, [pii] 10.1038/nsmb.2373 [DOI] [PubMed] [Google Scholar]
- 86.Grewal SI, Jia S (2007) Heterochromatin revisited. Nat Rev Genet 8(1):35–46. doi: nrg2008, [pii] 10.1038/nrg2008 [DOI] [PubMed] [Google Scholar]
- 87.Hahn MA, Wu X, Li AX et al. (2011) Relationship between gene body DNA methylation and intragenic H3K9me3 and H3K36me3 chromatin marks. PLoS One 6(4):e18844. doi: 10.1371/journal.pone.0018844 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Luco RF, Pan Q, Tominaga K et al. (2010) Regulation of alternative splicing by histone modifications. Science 327(5968):996–1000. doi: 10.1126/science.1184208 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Pradeepa MM, Sutherland HG, Ule J et al. (2012) Psip1/Ledgf p52 binds methylated histone H3K36 and splicing factors and contributes to the regulation of alternative splicing. PLoS Genet 8(5):e1002717. doi: 10.1371/journal.pgen.1002717 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Sims RJ 3rd, Millhouse S, Chen CF et al. (2007) Recognition of trimethylated histone H3 lysine 4 facilitates the recruitment of transcription postinitiation factors and pre-mRNA splicing. Mol Cell 28(4):665–676. doi: 10.1016/j.molcel.2007.11.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Gunderson FQ, Johnson TL (2009) Acetylation by the transcriptional coactivator Gcn5 plays a novel role in co-transcriptional spliceosome assembly. PLoS Genet 5(10):e1000682. doi: 10.1371/journal.pgen.1000682 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Klose RJ, Bird AP (2006) Genomic DNA methylation: the mark and its mediators. Trends Biochem Sci 31(2):89–97. doi: S0968–0004(05)00352-X, [pii] 10.1016/j.tibs.2005.12.008 [DOI] [PubMed] [Google Scholar]
- 93.Choi JK (2010) Contrasting chromatin organization of CpG islands and exons in the human genome. Genome Biol 11(7):R70. doi: gb-2010–11-7-r70, [pii] 10.1186/gb-2010-11-7-r70 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Feng S, Cokus SJ, Zhang X et al. (2010) Conservation and divergence of methylation patterning in plants and animals. Proc Natl Acad Sci USA 107(19):8689–8694. doi: 1002720107, [pii] 10.1073/pnas.1002720107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Zemach A, McDaniel IE, Silva P et al. (2010) Genome-wide evolutionary analysis of eukaryotic DNA methylation. Science 328(5980):916–919. doi: science.1186366, [pii] 10.1126/science.1186366 [DOI] [PubMed] [Google Scholar]
- 96.Lyko F, Foret S, Kucharski R et al. (2010) The honey bee epigenomes: differential methylation of brain DNA in queens and workers. PLoS Biol 8(11):e1000506. doi: 10.1371/journal.pbio.1000506 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Park J, Peng Z, Zeng J et al. (2011) Comparative analyses of DNA methylation and sequence evolution using Nasonia genomes. Mol Biol Evol 28(12):3345–3354. doi: msr168, [pii] 10.1093/molbev/msr168 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Bonasio R, Li Q, Lian J et al. (2012) Genome-wide and caste-specific DNA methylomes of the ants camponotus floridanus and harpegnathos saltator. Curr Biol 22(19):1755–1764. doi: S0960–9822(12)00867–6, [pii] 10.1016/j.cub.2012.07.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Flores KB, Wolschin F, Allen AN et al. (2012) Genome-wide association between DNA methylation and alternative splicing in an invertebrate. BMC Genomics 13(1):480. doi: 10.1186/1471-2164-13-480 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Lister R, Pelizzola M, Dowen RH et al. (2009) Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 462(7271):315–322. doi: nature08514, [pii] 10.1038/nature08514 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Hodges E, Smith AD, Kendall J et al. (2009) High definition profiling of mammalian DNA methylation by array capture and single molecule bisulfite sequencing. Genome Res 19(9):1593–1605. doi: gr.095190.109, [pii] 10.1101/gr.095190.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Laurent L, Wong E, Li G et al. (2010) Dynamic changes in the human methylome during differentiation. Genome Res 20(3):320–331. doi: gr.101907.109, [pii] 10.1101/gr.101907.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Pastor WA, Pape UJ, Huang Y et al. (2011) Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature 473(7347):394–397. doi: nature10102, [pii] 10.1038/nature10102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Ficz G, Branco MR, Seisenberger S et al. (2011) Dynamic regulation of 5-hydroxymethylcytosine in mouse ES cells and during differentiation. Nature 473(7347):398–402. doi: nature10008, [pii] 10.1038/nature10008 [DOI] [PubMed] [Google Scholar]
- 105.Khare T, Pai S, Koncevicius K et al. (2012) 5-hmC in the brain is abundant in synaptic genes and shows differences at the exon-intron boundary. Nat Struct Mol Biol 19(10):1037–1043. doi: 10.1038/nsmb.2372 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Shukla S, Kavak E, Gregory M et al. (2011) CTCF-promoted RNA polymerase II pausing links DNA methylation to splicing. Nature 479(7371):74–79. doi: 10.1038/nature10442 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Gowher H, Brick K, Camerini-Otero RD et al. (2012) Vezf1 protein binding sites genome-wide are associated with pausing of elongating RNA polymerase II. Proc Natl Acad Sci USA 109(7):2370–2375. doi: 10.1073/pnas.1121538109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Dickson J, Gowher H, Strogantsev R et al. (2010) VEZF1 elements mediate protection from DNA methylation. PLoS Genet 6(1):e1000804. doi: 10.1371/journal.pgen.1000804 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Kucharski R, Maleszka J, Foret S et al. (2008) Nutritional control of reproductive status in honeybees via DNA methylation. Science 319(5871):1827–1830. doi: 1153069, [pii] 10.1126/science.1153069 [DOI] [PubMed] [Google Scholar]
- 110.Oberdoerffer S (2012) A conserved role for intragenic DNA methylation in alternative pre-mRNA splicing. Transcription 3(3):106–109. doi: 19816, [pii] 10.4161/trns.19816 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Kim S, Kim H, Fong N et al. (2011) Pre-mRNA splicing is a determinant of histone H3K36 methylation. Proc Natl Acad Sci USA 108(33):13564–13569. doi: 10.1073/pnas.1109475108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.de Almeida SF, Grosso AR, Koch F et al. (2011) Splicing enhances recruitment of methyltransferase HYPB/Setd2 and methylation of histone H3 Lys36. Nat Struct Mol Biol 18(9):977–983. doi: 10.1038/nsmb.2123 [DOI] [PubMed] [Google Scholar]
- 113.Zhou HL, Hinman MN, Barron VA et al. (2011) Hu proteins regulate alternative splicing by inducing localized histone hyperacetylation in an RNA-dependent manner. Proc Natl Acad Sci USA 108(36):E627–E635. doi: 10.1073/pnas.1103344108 [DOI] [PMC free article] [PubMed] [Google Scholar]